Organocatalytic Asymmetric Biginelli-like Reaction Involving Isatin

Feb 2, 2016 - Dipartimento di Chimica, Materiali ed Ing. Chimica “Giulio Natta”, Politecnico di Milano, p.zza Leonardo da Vinci 32, 20133, Milano,...
1 downloads 0 Views 744KB Size
Subscriber access provided by La Trobe University Library

Article

Organocatalytic Asymmetric Biginelli-like Reaction Involving Isatin Mattia Stucchi, Giordano Lesma, Fiorella Meneghetti, Giulia Rainoldi, Alessandro Sacchetti, and Alessandra Silvani J. Org. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.joc.5b02680 • Publication Date (Web): 02 Feb 2016 Downloaded from http://pubs.acs.org on February 4, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Organic Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

Organocatalytic Asymmetric Biginelli-like Reaction Involving Isatin

Mattia Stucchi,a Giordano Lesma,a Fiorella Meneghetti,b Giulia Rainoldi,a Alessandro Sacchettic and Alessandra Silvania,* a

Università di Milano, Dipartimento di Chimica, via Golgi 19, Milano, 20133, Italy. bDipartimento di

Scienze Farmaceutiche, Università degli Studi di Milano, via L. Mangiagalli 25, 20133 Milano, Italy. c

Politecnico di Milano, Dipartimento di Chimica, Materiali ed Ing.Chimica ‘Giulio Natta’, p.zza Leonardo da Vinci 32, Milano, 20133, Italy.

ABSTRACT The first asymmetric, Brønsted acid-catalyzed Biginelli-like reaction of a ketone has been developed, employing N-substituted isatins as carbonyl substrates, and urea and alkyl acetoacetates as further components. BINOL-derived phosphoric acid catalysts have been used to achieve the synthesis of a small library of chiral, enantioenriched spiro(indoline-pyrimidine)-diones derivatives. The absolute configuration of the new spiro stereocenter was assessed on diastereoisomeric derivatives through computer-assisted NMR spectroscopy. X-Ray diffractometry allowed to disclose the overall molecular conformation in the solid state and to characterize the crystal packing of a Br-substituted Biginelli-like derivative, while computational studies on the reaction transition state allowed to rationalize the stereochemical outcome.

INTRODUCTION 2-Oxindoles, especially those 3,3-disubstituted or spiro-fused to other cyclic frameworks, continue to be recognized as valuable compounds for drug discovery. They feature in a large number of natural and

ACS Paragon Plus Environment

The Journal of Organic Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 25

unnatural compounds with important biological activities and serve as key intermediates for the synthesis of many kinds of drug candidates.1 In particular, spirooxindoles, having cyclic structures fused at the C3 carbon, move away from the flat heterocycles encountered in many drug discovery programs. For this reason, they are of special interest, being able to potentially provide improved physicochemical properties in their interaction with biological systems.2 As more examples of the enantiospecific biological activity are identified, efficient and reliable asymmetric synthesis of such compounds becomes ever more valuable. In particular, the improvement of practical and versatile multicomponent approaches has attracted considerable interest owing to their synthetic efficiency and extensive diversity-generating ability.3 Multicomponent reactions (MCRs) are very efficient tools to quickly prepare pharmacological compounds. However, their combination with asymmetric catalysis, in particular organocatalysis, remains a largely unmined area of research, although the results reported until now show the possibilities and versatility of this type of strategy, which allows elevate levels of atom efficiency and enantioselectivity to be reached at the same time.4 In the field of oxindole chemistry, to date only a few organocatalyzed multicomponent methods have been reported towards the asymmetric generation of the structurally rigid architecture of 3,3-disubstituted or spiro-fused oxindoles.5 Noteworthy among them, is the cinchona alkaloid derived amine-catalyzed Michael-type addition developed in highly efficient three-component versions using readily available malononitrile, isatins and ketones.6,7 Quite recently, isatin-derived 3-indolylmethanols have emerged as useful substrates for phosphoric acid-catalyzed three component cascade Michael/Pictet-Spengler reactions.8 On the other hand, intense effort have been devoted to develop organocatalytic MCRs to form spiro[pyrrolidin-3,2'-oxindoles] and spirooxindole pyran derivatives by means of 1,3-dipolar cycloadditions9 or cascade [3+2]10 or [2+2+2] cycloadditions.11 As part of our interest in the asymmetric synthesis of 3,3-disubstituted oxindole derivatives and related spiro-compounds,12 we turned our attention to the MCRs field, in order to explore the single reactant ACS Paragon Plus Environment

Page 3 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

replacement (SRR) approach.13 By this strategy, starting from a well-known MCR, new applications can be found, just replacing a single component with a different input able to carry out the key chemical reactivity necessary for that MCR to occur. In this context, we focused on the Biginelli reaction, one of the well established MCRs, mainly employed for the synthesis of 3,4-dihydropyrimidine-2(1H)-ones (DHPMs). Such heterocyclic scaffolds have found increasing applications in medicinal chemistry, because of their important pharmacological and biological properties.14 Only few examples are reported on enantioselective organocatalytic Biginelli reactions, all involving aromatic aldehydes as carbonyl components.15 The milestone was placed by Gong,16,17,18 who disclosed the first highly enantioselective protocol, based on BINOL-derived chiral phosphoric acids as organocatalysts. Also dual-activation routes have been developed, by using combined catalysts consisting of a Brønsted acid and a chiral secondary amine19,20 or, alternatively, a chiral bifunctional primary amine-thiourea.21 At the best of our knowledge, only two examples of the multicomponent preparation of racemic DHPMs derivatives starting from isatin are reported.22,23 In general, application of organocatalysis to the Biginellilike reaction, employing a ketone as the carbonyl component, is even now quite unexplored. Herein, we report the Brønsted acid-catalyzed asymmetric synthesis of spiro(indoline-pyrimidine)-diones derivatives via a Biginelli-like reaction, consisting of a three component cyclocondensation of alkyl acetoacetates, urea and isatin derivatives instead of aldehydes (Scheme 1).

ACS Paragon Plus Environment

The Journal of Organic Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 25

Scheme 1. Strategy used for the asymmetric construction of the spiro(indoline-pyrimidine)-dione scaffold.

RESULTS AND DISCUSSION Our initial studies were performed taking into account the Brønsted acid-catalytic enantioselective protocol reported by Gong for the true, aldehyde-involving, Biginelli reaction. Isatin 1a, urea 2, ethyl acetoacetate 3 and (R)-BINOL-derived phosphoric acid 4a were chosen for preliminary experiments (Table 1). Table 1. Optimization of the asymmetric Biginelli-like reaction.a

entry

catalyst

solvent

1 2 3 4 5 6 7 8 9 10 11

4a 4a 4a 4a 4b 4c 4d 4e 4f 4g 4a

CH2Cl2 Toluene CH2Cl2 Toluene Toluene Toluene Toluene Toluene Toluene Toluene Toluene

conc. [mol/L] 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2

temp. [°C], time [h] rt, 96 rt, 96 50, 48 50, 48 50, 48 50, 48 50, 48 50, 48 50, 48 50, 48 50, 96

yieldb [%] trace trace 32 51 trace 30 31 trace trace 26 60

eec [%] / / 75 80 / 79 81 / / 79 80

ACS Paragon Plus Environment

Page 5 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

12 4a 13 4a 14 4a a Reactions were

The Journal of Organic Chemistry

Toluene Toluene Toluene performed on

0.2 50, 240 0.2 70, 96 0.4 50, 96 a 0.16 mmol scale with

66 77 65 48 62 66 1/2/3 in 1/1.2/3 ratio, in the presence of 20 mol %

(R)-4 (0.032 mmol). bIsolated yield. cDetermined by chiral HPLC analysis.

At room temperature, the reaction proceeds with difficulty both in CH2Cl2 and in toluene (entries 1 and 2) and, after 96 hours, only trace amounts of the desired compound 5a could be detected by 1H NMR of the crude reaction mixture. The lower reactivity of C-3 carbonyl group of isatin compared to aldehydes, along with its higher steric demand, appears to be the key factor hindering the reaction from successfully proceeding at room temperature. To our delight, increasing the temperature to 50 °C (entries 3 and 4) entailed a significant effect on the chemical conversion. Toluene proved to be the solvent of choice, affording product 5a in acceptable yield and with a good level of enantioselectivity. Screening of more hindered catalysts 4b-f, aimed to evaluate the impact of the 3,3'-substitution, and of octahydro-BINOL-based 4g was performed (entries 5-10). Increasing the size of the 3,3'-substitents on the phosphoric acid proved detrimental for the chemical conversion, with only catalysts 4c and 4d able to afford product 5a, with maintenance of the same level of enantioselectivity as 4a, but in definitely decreased yields. After that, we established 4a as the catalyst of choice, and further screening of the reaction conditions were performed. Some yield improvement without sacrificing the stereoselectivity could be achieved by prolonging the reaction time until 96 hours (entry 11). More prolonged times are not convenient for the balance among yield and ee (entry 12). Increasing the reaction temperature deeply eroded the enantioselectivity, albeit with better yield (entry 13). The same happened when the reaction was conducted in more concentrated conditions (entry 14). Lowering the reactant concentration or the catalyst loading led to a significant decrease in yield.

ACS Paragon Plus Environment

The Journal of Organic Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 25

After establishing the optimal conditions, the Biginelli-like reaction of a isatins series was examined, using (R)-4a as catalyst, in toluene at 50 °C for 96 h (Figure 1).

H N

O

H N

O HN

HN

CO2Et O

H N

O HN

CO2Et O

CO2Et O

HN

N

N

N

H N

O

CO2Et O N

O2N

MeO 5a (60%, ee: 80%) H N

O HN

F

5b (63%, ee: 75%) H N

O CO2Et O

HN

Cl

O

H N

O CO2Me O

N

5i (65%, ee: 61%)

Cl

5d (93%, ee: 50%)

H N

HN

CO2Et O

5f (66%, ee: 74%)

O

N

Br

5g (62%, ee: 77%)

CO2Bn O N

5j (55%, ee: 74%)

Figure 1. Substrate scope of the Biginelli-like reaction catalyzed by (R)-4a.

ACS Paragon Plus Environment

H N

HN

CO2Et O

H N

HN

HN

O

N

N

5e (51%, ee: 75%)

5c (49%, ee: 70%)

CO2Et O N

5h (63%, ee: 75%)

Page 7 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

The substrate scope was surveyed, by evaluating differently N-substituted isatins and the presence of substituents at 5- or 6-position of the isatin nucleus. In general, all isatins readily undergo this reaction, to afford the desired products 5a-h in moderate to high yields, with a good degree of enantioselectivity. Only the sterically demanding N-trityl isatin failed to participate in the reaction and the corresponding Biginellilike adduct could not be detected. The N-Me isatin gave a better result than the corresponding N-Benzyl, Np-Nitrobenzyl and N-p-Methoxybenzyl ones in terms of yield (93% in comparison to up to 63%), but suffering a drop in ee (50% in comparison to up to 80%). The presence of various halogen substituents at the aryl ring has almost no effect on both yield and ee. Variations at the ester moiety of the β-ketoester component were also evaluated. Methyl and benzyl acetoacetates participated at the reaction efficiently to provide adducts 5i-j in good yields and moderate ee's. In this kind of reaction, surprisingly, neither thiourea in place of urea, nor various linear or cyclic βdiketones in place of alkyl acetoacetates, showed to be suitable, together with N-benzyl-isatin. With thiourea no reaction occurred, while with β-diketones a complex mixture of products could be detected. Then, we examined some products transformations, first of all the facile regioselective mono-N-alkylation of the dihydropyrimidin-2-one ring. Starting from the Biginelli-like compound 5a, the corresponding Nbenzyl derivative 6 was achieved in high yield and regioselectivity, by reaction with benzyl bromide and cesium carbonate, in DMF at room temperature (Scheme 2). Further, catalytic hydrogenolysis of the benzyl ester moiety of compound 5j allowed to easily obtain the carboxylic acid derivative 7, that can be regarded as a useful key intermediate toward the synthesis of peptidomimetic compounds. The carboxylic acid functional group of 7 can also be quantitatively removed to give 8, by heating in acidic conditions.

ACS Paragon Plus Environment

The Journal of Organic Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 25

Scheme 2. Synthetic transformations on compounds 5a and 5j.

In order to demonstrate the reactivity of acid 7 and aiming at the same time to gain information on the absolute configuration of the major enantiomer 5j (vide infra), obtained in the (R)-4a-catalyzed Biginellilike reaction, we pursued the transformation depicted in Scheme 3.

O

H N

HN

CO2H + O

NH2

i) HATU, DIPEA, DMF, rt ii) chromatographic column separation

N Bn

98%

9a:9b 87:13

7 (from 5j, ee: 74%)

H N

O

O H N 1'

HN 3

+

H N H N 1'

HN

OO

N Bn

3

OO

N Bn 9a

9b

Scheme 3. Synthesis of diastereoisomeric compounds 9a and 9b, starting from acid 7.

ACS Paragon Plus Environment

Page 9 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

By reaction with (S)-1-phenylethanamine in the presence of the condensing agent HATU, acid 7 was cleanly converted into diastereoisomeric amides 9a and 9b, which could be efficiently separated by flash chromatography, establishing the possible application of 7 in peptidomimetic chemistry. Confiding at first on X-ray diffractometry in order to determine the C3 absolute configuration of compounds 5, we planned to perform the crystallographic analysis on 5h. This molecule was selected as a suitable derivative, due to the presence of the bromine atom as anomalous dispersor. Initially, 5h disclosed a recalcitrant crystallization behavior in yielding single crystals and, only after many attempts, well diffracting crystals were obtained. The X-ray data revealed that the 12:88 molar mixture of enantiomers crystallized in a centrosymmetric space group, showing the more favored crystallization of the racemate instead of the major enantiomer. In the solid state, the overall molecular conformation is determined by the spiro(indoline-pyrimidine)-dione system, with the dihydropyrimidin-2-one ring, having an almost planar conformation, perpendicularly oriented with respect to oxindole (Figure 2a). The conformation of the benzyl group shows the phenyl ring pointing in the same direction of the dihydropyrimidin-2-one carbonyl moiety (see Supporting Information). The crystal packing is characterized by strong centrosymmetric NH...O hydrogen bonds, leading to the formation of dimers, that are in turns stabilized by Cπ-H...O contacts, as depicted in Figure 2b. This interactions pattern can be employed for rationalizing the preferential crystallization of the racemate, which is indeed consistent with the close packing found in the crystal environment, dominated by unique characteristics of hydrogen bonds involved in dimers formation. This easier racemate crystallization is in agreement with previous literature data,24 showing the tendency for several racemic crystals to be more stable and denser than their chiral counterparts.

ACS Paragon Plus Environment

The Journal of Organic Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 25

Figure 2. (a) ORTEP25 drawing of 5h, showing the arbitrary atomic numbering (displacement ellipsoids at 40% probability). (b) Intermolecular interactions viewed along c axis.

Although it was not possible to obtain suitable crystals for X-ray-based determination of the prevailing enantiomer 5h, we were able to determine the C3 stereochemistry through ab initio calculation of NMR shifts, a technique pioneered by Bifulco.26 We considered the differences in both 1H and 13C NMR spectra for compounds 9a and 9b and then performed a theoretical conformational search on both (3S,1’S) and (3R,1’S) possible diastereoisomers, employing the Monte Carlo algorithm and molecular mechanics (MMFF force field). After DFT optimization, we calculated 1H and

13

C NMR chemical shifts, by

subjecting the shielding constants to Boltzmann averaging over the conformers, followed by linear regression, as reported by Pierens.27 From comparison of experimental and calculated data, the (3S,1'S) absolute configuration could be confidently assigned to the major diastereoisomer 9a and, consequently, the (3R,1'S) one to the minor 9b. To make this assignment safe beyond any doubt, we also calculated the comparison parameter (CP3), especially designed28 for the computer-assisted assignment of the stereochemistry of diastereoisomers pairs, in which only the configuration of one stereocenter is unknown. ACS Paragon Plus Environment

Page 11 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

By this way, our stereochemical assignment could be made quite secure also from a quantitative point of view (see Supporting Information). These results allowed us to disclose the C3-S favoring enantioselectivity of the described organocatalyzed reaction and prompted us to perform theoretical calculations on the stereogenic centre forming step. The mechanism of the Biginelli reaction has been previously investigated by means of computational tools,29 also in the presence of tartaric acid as catalyst.30 Results indicated the iminium path as the most favorable, in accordance with previously proposed mechanism.31 Therefore, we decided to investigate the initial addition of the enol form of ethyl acetoacetate on the imine formed between isatin 1a and urea 2, in the presence of (R)-4a, since in this step the final configuration of 5a is determined. DFT study at the B3LYP/6-31G(d,p) level of theory were performed taking into account the two possible spatial arrangement of the more stable Z-imine in the reagents-catalyst complex,32 leading to the diastereoisomeric transition state models TS-A and TS-B (Figure 3). All the calculations were performed with the Spartan ’08 33 suite (see Supporting Information). The energy profiles clearly indicate a strong preference for TS-A, with a ∆∆G‡ = 1.47 Kcal/mol with respect to TS-B, at T = 323K, from which an expected 85% ee could be calculated. These results are in satisfactory agreement with the experimental observed ee's, and once again support the previously predicted S configuration for major diastereoisomer 9a.

ACS Paragon Plus Environment

The Journal of Organic Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 25

Figure 3. Proposed transition states TS-A and TS-B ( and the corresponding 3D structures) of the BINOLderived phosphoric acid catalyzed Biginelli-like reaction to give 5a. In 3D TS-B red lines highlight the steric hindrance between one phenyl substituent of (R)-4a and the ureidic residue.

Looking at the transition states 3D structures, the steric hindrance between one phenyl substituent of (R)-4a and the ureidic residue could explain the higher activation energy of TS-B and the resulting favored nucleophilic attack on the si-face of the imine (TS-A). Moreover, in TS-A an hydrogen bond between the ureidic NH of the imine and the carbonyl oxygen of the acetoacetate ester is established, thus further stabilizing this structure. CONCLUSION In conclusion, we developed the first enantioselective organocatalyzed Biginelli-like reaction applied to a ketone, namely isatin, with good yields and enantioselectivity. By employing BINOL-based phosphoric acids as catalysts and different isatins and alkyl acetoacetates as substrates, together with urea, a small library of enantioenriched spiro[indoline-pyrimidine]-dione derivatives could be obtained. Postcondensation reactions have been performed, increasing the number of potentially useful compounds. ACS Paragon Plus Environment

Page 13 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

The solid state conformation of a Br-containing Biginelli-like compound was investigated, putting in evidence its crystallization behaviour leading to the more favored racemate, instead of the major enantiomer. The absolute configuration at the oxindole C3 quaternary stereocenter was assessed to be S for the major enantiomer, by means of quantum mechanical methods and NMR spectroscopy on diastereoisomeric derivatives. Computational studies on the reaction transition state (TS) allowed us to explain the experimentally observed enantioselectivity and stereochemical outcome.

EXPERIMENTAL SECTION General information All commercial materials were used without further purification. All solvents were of reagent grade or HPLC grade. All reactions were carried out under a nitrogen atmosphere unless otherwise noted. All reactions were monitored by thin layer chromatography (TLC) on precoated silica gel 60 F254; spots were visualized with UV light or by treatment with a 1% aqueous KMnO4 solution. Products were purified by flash chromatography on silica gel 60 (230–400 mesh). 1H NMR spectra and 13C NMR spectra were recorded on 300 and 400 MHz spectrometers. Chemical shifts are reported in parts per million relative to the residual solvent. 13C NMR spectra have been recorded using the APT pulse sequence. Multiplicities in 1H NMR are reported as follows: s = singlet, d = doublet, t = triplet, m = multiplet, br s = broad singlet. High-resolution MS spectra were recorded with a Q-ToF mass spectrometer, equipped with an ESI source. Chiral HPLC analysis was performed with UV Detector and binary HPLC pump at 254 nm. Chiralcel OD column was used. Specific optical rotation [α]TD was measured with a cell of 1 dm pathlength and 1 mL capacity. The light used has a wavelength of 589 nm (Sodium D line). N-substituted isatins34 and BINOLphosphoric acids35 were synthetized according to reported literature.

General procedure for the asymmetric organocatalyzed synthesis of compounds 5a-j ACS Paragon Plus Environment

The Journal of Organic Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 25

Substituted isatin 1 (0.16 mmol, 1 eq), urea 2 (0.19 mmol, 1.2 eq), alkyl acetoacetate 3 (0.48 mmol, 3 eq) and (R)-4a catalyst (0.03 mmol, 0.2 eq) were dissolved in toluene (0.800 mL, 0.2 M). The reaction was stirrer at 50 °C for 96 hours. The resulting mixture was then concentrated under reduced pressure, to give a residue which was purified by flash chromatography (FC) as indicated below. (S)-ethyl 1-benzyl-6'-methyl-2,2'-dioxo-2',3'-dihydro-1'H-spiro[indoline-3,4'-pyrimidine]-5'-carboxylate 5a Prepared according to general procedure starting from N-benzyl isatin and ethyl acetoacetate; FC: dichlorometane:methanol, 97.5:2.5; yield: 60%; white solid; mp 223-224 oC; [α]20D – 45.5 (c 0.2, CHCl3); 1

H NMR (400 MHz, CDCl3) δ 8.50 (br s, 1H), 7.42 (d, J = 7.4 Hz, 2H), 7.38 – 7.24 (m, 4H), 7.21 (t, J = 7.5

Hz, 1H), 7.03 (t, J = 7.5 Hz, 1H), 6.76 (d, J = 7.8 Hz, 1H), 5.69 (br s, 1H), 4.99 (d, J = 15.5 Hz, 1H), 4.80 (d, J = 15.5 Hz, 1H), 3.99 – 3.86 (m, 1H), 3.70 – 3.55 (m, 1H), 2.38 (s, 3H), 0.71 (t, J = 7.1 Hz, 3H). 13C NMR (100 MHz, CDCl3) δ 176.5, 165.2, 151.9, 149.9, 143.2, 136.3, 132.9, 130.5, 129.5 (2C), 128.5 (3C), 124.6, 124.0, 109.9, 99.4, 64.2, 60.6, 45.0, 20.1, 14.1; HRMS (ESI) calcd for C22H21N3NaO4+ [MNa]+ 414.1434, found 414.1442; enantiomeric excess: 80%, determined by chiral HPLC (n-hexane:isopropanol = 80:20, flow rate 1.0 mL/min): tR = 14.98 min (major), tR = 33.78 min (minor). (S)-ethyl

1-(4-methoxybenzyl)-6'-methyl-2,2'-dioxo-2',3'-dihydro-1'H-spiro[indoline-3,4'-pyrimidine]-5'-

carboxylate 5b Prepared according to general procedure starting from N-(4-methoxybenzyl) isatin and ethyl acetoacetate; FC: dichlorometane:methanol, 97.5:2.5; yield: 63%; white solid; mp 193-194 oC; [α]20D + 4.5 (c 0.35, CHCl3); 1H NMR (300 MHz, CDCl3, mixture of conformers 6:1) δ 8.87 (br s, 0.15H), 8.74 (br s, 0.85H), 7.33 (d, J = 8.7 Hz, 2H), 7.26 (d, J = 8.2 Hz, 1H), 7.17 (t, J = 7.7 Hz, 1H), 6.98 (t, J = 7.3 Hz, 1H), 6.88 – 6.70 (m, 3H), 6.01 (br s, 0.86H), 5.81 (br s, 0.14H), 4.85 (d, J = 15.3 Hz, 1H), 4.71 (d, J = 15.3 Hz, 1H), 3.96 – 3.78 (m, 1H), 3.74 (s, 0.43H), 3.71 (s, 2.57H), 3.64 – 3.43 (m, 1H), 2.33 (s, 0.44H), 2.27 (s, 2.56H), 0.64 (t, J = 7.1 Hz, 3H).

13

C NMR (75 MHz, CDCl3, mixture of conformers 6:1) δ 175.9, 164.6, 159.1,

152.1, 149.5, 142.6, 132.5, 129.7, 129.2 (2C), 127.8, 123.9, 123.2, 114.1 (2C), 109.2, 98.5, 63.4, 59.8, 55.2, ACS Paragon Plus Environment

Page 15 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

43.7, 19.1, 13.5; HRMS (ESI) calcd for C23H23N3NaO5+ [MNa]+ 444.1530, found 444.1519; enantiomeric excess: 75%, determined by chiral HPLC (n-hexane:isopropanol = 65:35, flow rate 1.0 mL/min): tR = 9.85 min (major), tR = 27.96 min (minor). (S)-ethyl

6'-methyl-1-(4-nitrobenzyl)-2,2'-dioxo-2',3'-dihydro-1'H-spiro[indoline-3,4'-pyrimidine]-5'-

carboxylate 5c Prepared according to general procedure starting from N-(4-nitrobenzyl) isatin and ethyl acetoacetate; FC: dichlorometane:methanol, 97.5:2.5; yield: 49%; white solid; mp 201-202 oC; [α]20D – 8.2 (c 0.3, CHCl3); 1H NMR (300 MHz, CDCl3, mixture of conformers 5:1) δ 8.48 (br s, 0.17H), 8.40 (br s, 0.83H), 8.21 – 8.08 (m, 2H), 7.65 – 7.54 (m, 2H), 7.30 (d, J = 7.3 Hz, 1H), 7.19 (t, J = 7.3 Hz, 1H), 7.03 (t, J = 7.5 Hz, 1H), 6.62 (d, J = 7.6 Hz, 1H), 6.28 (br s, 0.84H), 6.12 (br s, 0.16H), 5.09 – 4.87 (m, 2H), 4.07 – 3.89 (m, 1H), 3.86 – 3.68 (m, 1H), 2.34 (s, 0.5H), 2.30 (s, 2.5H), 0.88 (m, 3H).

13

C NMR (75 MHz, CDCl3, mixture of

conformers 5:1) δ 176.0, 164.6, 151.8, 149.0, 147.5, 143.0, 141.9, 132.2, 130.0, 128.4 (2C), 124.1, 124.0 (2C), 123.8, 109.0, 98.9, 63.51, 60.3, 43.7, 19.5, 13.8; HRMS (ESI) calcd for C22H20N4NaO6+ [MNa]+ 459.1275, found 459.1268; enantiomeric excess: 70%, determined by chiral HPLC (n-hexane:isopropanol = 50:50, flow rate 1.0 mL/min): tR = 10.05 min (major), tR = 45.50 min (minor). (S)-ethyl 1,6'-dimethyl-2,2'-dioxo-2',3'-dihydro-1'H-spiro[indoline-3,4'-pyrimidine]-5'-carboxylate 5d Prepared according to general procedure starting from N-methyl isatin and ethyl acetoacetate; FC: dichlorometane:methanol, 95:5; yield: 93%; white solid; mp 228-229 oC; [α]20D – 1.6 (c 0.35, CHCl3); 1H NMR (400 MHz, DMSO-d6) δ 9.45 (br s, 1H), 7.75 (br s, 1H), 7.30 (t, J = 7.7 Hz, 1H), 7.18 (d, J = 7.3 Hz, 1H), 7.01 (t, J = 7.4 Hz, 1H), 6.97 (d, J = 7.8 Hz, 1H), 3.68 (q, J = 7.1 Hz, 2H), 3.10 (s, 3H), 2.26 (s, 3H), 0.75 (t, J = 7.1 Hz, 3H). 13C NMR (100 MHz, DMSO-d6) δ 176.2, 164.8, 150.9 (2C), 144.0, 134.0, 129.6, 123.4, 122.8, 108.7, 97.1, 63.1, 59.5, 26.6, 18.7, 13.8; HRMS (ESI) calcd for C16H17N3NaO4+ [MNa]+ 338.1111, found 338.1123; enantiomeric excess: 50%, determined by chiral HPLC (n-hexane:isopropanol = 65:35, flow rate 1.0 mL/min): tR = 7.25 min (major), tR = 38.20 min (minor). ACS Paragon Plus Environment

The Journal of Organic Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(S)-ethyl

Page 16 of 25

1-benzyl-5-fluoro-6'-methyl-2,2'-dioxo-2',3'-dihydro-1'H-spiro[indoline-3,4'-pyrimidine]-5'-

carboxylate 5e Prepared according to general procedure starting from 5-fluoro-N-benzyl isatin and ethyl acetoacetate; FC: dichlorometane:methanol, 97.5:2.5; yield: 51%; white solid; mp 135-136 oC; [α]20D + 3.8 (c 0.3, CHCl3); 1H NMR (300 MHz, CDCl3, mixture of conformers 5:1) δ 8.76 – 8.49 (br, m, 1H), 7.38 (d, J = 7.2 Hz, 2H), 7.34 – 7.16 (m, 3H), 7.03 (dd, J = 7.3, 2.5 Hz, 1H), 6.88 (td, J = 8.8, 2.4 Hz, 1H), 6.67 (dd, J = 8.5, 3.8 Hz, 1H), 6.13 – 5.88 (br, m, 1H), 4.92 (d, J = 15.5 Hz, 1H), 4.76 (d, J = 15.5 Hz, 1H), 4.04 – 3.83 (m, 1H), 3.74 – 3.52 (m, 1H), 2.35 (s, 0.5H), 2.30 (s, 2.5H), 0.83 – 0.68 (m, 3H). 13C NMR (75 MHz, CDCl3, mixture of conformers 5:1) δ 175.7, 164.4, 161.1, 157.9, 151.80, 149.75, 138.42, 135.34, 128.77 (2C), 127.81, 127.74 (2C), 116.17 and 115.86 (1C), 112.19 and 111.9 (1C), 110.0 and 109.9 (1C), 98.2, 63.6, 60.1, 44.4, 19.3, 13.6; HRMS (ESI) calcd for C22H20FN3NaO4+ [MNa]+ 432.1330, found 432.1326; enantiomeric excess: 75%, determined by chiral HPLC (n-hexane:isopropanol = 70:30, flow rate 1.0 mL/min): tR = 8.15 min (major), tR = 16.35 min (minor). (S)-ethyl

1-benzyl-5-chloro-6'-methyl-2,2'-dioxo-2',3'-dihydro-1'H-spiro[indoline-3,4'-pyrimidine]-5'-

carboxylate 5f Prepared according to general procedure starting from 5-chloro-N-benzyl isatin and ethyl acetoacetate; FC: dichlorometane:methanol, 97.5:2.5; yield: 66%; white solid; mp 126-127 oC; [α]20D + 33.6 (c 0.2, CHCl3); 1

H NMR (300 MHz, CDCl3, mixture of conformers 5:1) δ 8.76 (br s, 0.16H), 8.69 (br s, 0.84H), 7.37 (d, J

= 7.1 Hz, 2H), 7.33 – 7.18 (m, 4H), 7.14 (d, J = 8.3, 1H), 6.67 (d, J = 8.3 Hz, 1H), 6.24 (br s, 0.83H), 6.18 (br s, 0.17H), 4.90 (d, J = 15.6 Hz, 1H), 4.77 (d, J = 15.7 Hz, 1H), 4.01 – 3.84 (m, 1H), 3.75 – 3.56 (m, 1H), 2.34 (s, 0.5H), 2.29 (s, 2.5H), 0.83 – 0.68 (m, 3H). 13C NMR (75 MHz, CDCl3) δ 175.5, 164.4, 151.9, 149.8, 141.1, 135.2, 134.0, 129.7, 128.8 (2C), 128.5, 127.8, 127.7 (2C), 124.3, 110.3, 98.1, 63.4, 60.1, 44.4, 19.3, 13.6; HRMS (ESI) calcd for C22H20ClN3NaO4+ [MNa]+ 448.1035, found 448.1049; enantiomeric

ACS Paragon Plus Environment

Page 17 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

excess: 74%, determined by chiral HPLC (n-hexane:isopropanol = 70:30, flow rate 1.0 mL/min): tR = 9.05 min (major), tR = 16.45 min (minor). (S)-ethyl

1-benzyl-6-chloro-6'-methyl-2,2'-dioxo-2',3'-dihydro-1'H-spiro[indoline-3,4'-pyrimidine]-5'-

carboxylate 5g Prepared according to general procedure starting from 6-chloro-N-benzyl isatin and ethyl acetoacetate; FC: dichlorometane:methanol, 97.5:2.5; yield: 62%; white solid; mp 213-214 oC; [α]20D – 1.0 (c 0.3, CHCl3); 1H NMR (300 MHz, CDCl3) δ 8.44 (br s, 1H), 7.39 (d, J = 7.0 Hz, 2H), 7.35 – 7.22 (m, 3H), 7.19 (d, J = 7.7 Hz, 1H), 6.98 (d, J = 7.6 Hz, 1H), 6.79 (s, 1H), 6.14 (br s, 1H), 4.90 (d, J = 15.5 Hz, 1H), 4.75 (d, J = 15.5 Hz, 1H), 3.90 (m, 1H), 3.60 (m, 1H), 2.29 (s, 3H), 0.74 (t, J = 6.8 Hz, 3H). 13C NMR (75 MHz, CDCl3) δ 175.8, 164.4, 151.9, 149.4, 143.9, 135.6, 135.1, 130.7, 128.8 (2C), 127.9, 127.8 (2C), 124.8, 123.1, 109.9, 98.4, 63.0, 60.1, 44.4, 19.2, 13.6; HRMS (ESI) calcd for C22H20ClN3NaO4+ [MNa]+ 448.1035, found 448.1039; enantiomeric excess: 77%, determined by chiral HPLC (n-hexane:isopropanol = 65:35, flow rate 1.0 mL/min): tR = 8.65 min (major), tR = 15.35 min (minor). (S)-ethyl

1-benzyl-6-bromo-6'-methyl-2,2'-dioxo-2',3'-dihydro-1'H-spiro[indoline-3,4'-pyrimidine]-5'-

carboxylate 5h Prepared according to general procedure starting from 6-bromo-N-benzyl isatin and ethyl acetoacetate; FC: dichlorometane:methanol, 97.5:2.5; yield: 63%; white solid; mp 221-222 oC; [α]20D + 7.6 (c 0.55, CHCl3); 1

H NMR (300 MHz, CDCl3) δ 8.52 (br s, 1H), 7.39 (d, J = 7.3 Hz, 2H), 7.26 (m, 3H), 7.18 – 7.06 (m, 2H),

6.95 (s, 1H), 6.22 (br s, 1H), 4.87 (d, J = 15.6 Hz, 1H), 4.75 (d, J = 15.6 Hz, 1H), 4.00 – 3.77 (m, 1H), 3.70 – 3.49 (m, 1H), 2.27 (s, 3H), 0.72 (t, J = 7.1 Hz, 3H). 13C NMR (75 MHz, CDCl3) δ 175.7, 164.4, 152.0, 149.5, 144.0, 135.1, 131.3, 128.8 (2C), 127.9, 127.8 (2C), 126.1, 125.2, 123.4, 112.6, 98.3, 63.1, 60.1, 44.1, 19.2, 13.6; HRMS (ESI) calcd for C22H20BrN3NaO4+ [MNa]+ 492.0529, found 492.0518; enantiomeric excess: 75%, determined by chiral HPLC (n-hexane:isopropanol = 70:30, flow rate 1.0 mL/min): tR = 9.35 min (major), tR = 16.50 min (minor). ACS Paragon Plus Environment

The Journal of Organic Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(S)-methyl

Page 18 of 25

1-benzyl-6'-methyl-2,2'-dioxo-2',3'-dihydro-1'H-spiro[indoline-3,4'-pyrimidine]-5'-carboxylate

5i Prepared according to general procedure starting from N-benzyl isatin and methyl acetoacetate; FC: dichlorometane:methanol, 97.5:2.5; yield: 65%; white solid; mp 143-144 oC; [α]20D – 6.8 (c 0.35, CHCl3); 1

H NMR (300 MHz, CDCl3) δ 7.77 (br s, 1H), 7.40 (d, J = 7.7 Hz, 2H), 7.36 – 7.23 (m, 4H), 7.20 (td, J =

7.8, 1.2 Hz, 1H), 7.01 (t, J = 7.5 Hz, 1H), 6.75 (d, J = 7.7 Hz, 1H), 5.29 (br s, 1H), 4.94 (d, J = 15.5 Hz, 1H), 4.83 (d, J = 15.4 Hz, 1H), 3.20 (s, 3H), 2.37 (s, 3H).

13

C NMR (75 MHz, CDCl3) δ 175.8, 165.0,

151.3, 149.2, 142.3, 135.6, 132.1, 129.9, 128.8 (2C), 127.9 (2C), 127.8, 123.8, 123.3, 109.2, 98.7, 63.5, 51.0, 44.3, 19.4; HRMS (ESI) calcd for C21H19N3NaO4+ [MNa]+ 400.1268, found 400.1257; enantiomeric excess: 61%, determined by chiral HPLC (n-hexane:isopropanol = 70:30, flow rate 1.0 mL/min): tR = 9.15 min (major), tR = 18.30 min (minor). (S)-benzyl

1-benzyl-6'-methyl-2,2'-dioxo-2',3'-dihydro-1'H-spiro[indoline-3,4'-pyrimidine]-5'-carboxylate

5j Prepared according to general procedure starting from N-benzyl isatin and benzyl acetoacetate; FC: dichlorometane:methanol, 97.5:2.5; yield: 55%; white solid; mp 147-148 oC; [α]20D +17.2 (c 0.5, dioxane); 1

H NMR (300 MHz, CDCl3) δ 8.15 (br s, 1H), 7.41 – 7.17 (m, 10H), 7.18 – 7.09 (m, 1H), 6.98 (t, J = 7.5

Hz, 1H), 6.85 (d, J = 6.5 Hz, 1H), 6.44 (d, J = 7.8 Hz, 1H), 5.39 (br s, 1H), 4.85 – 4.70 (m, 2H), 4.63 (d, J = 12.0 Hz, 1H), 3.81 (d, J = 15.6 Hz, 1H), 2.39 (s, 3H). 13C NMR (75 MHz, CDCl3) δ 175.8, 164.4, 162.2, 150.2, 142.3, 135.8, 132.2, 129.9, 128.9 (4C), 128.6 (2C), 128.3, 127.8, 127.7 (2C), 124.0, 123.4, 109.8, 66.5, 43.8, 19.6 (3 quaternary carbons are missed); HRMS (ESI) calcd for C27H23N3NaO4+ [MNa]+ 476.1581, found 476.1589; enantiomeric excess: 74%, determined by chiral HPLC (n-hexane:isopropanol = 80:20, flow rate 1.0 mL/min): tR = 13.50 min (major), tR = 28.20 min (minor). Procedure for the synthesis of ethyl (S)-1,1'-dibenzyl-6'-methyl-2,2'-dioxo-2',3'-dihydro-1'Hspiro[indoline-3,4'-pyrimidine]-5'-carboxylate (6) ACS Paragon Plus Environment

Page 19 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

To a solution of compound 5a (0.25 mmol, 1 eq) in anhydrous dimethylformamide (0.830 mL, 0.3 M), CsCO3 (0.33 mmol, 1.3 eq) was added, then the mixture was stirred for 1 hour at room temperature. Benzyl bromide (0.38 mmol, 1.5 eq) was slowly added and the mixture was stirred overnight. After the completion of reaction (monitored by TLC), saturated aq. NaCl (1 mL) was added. The reaction mixture was extracted with ethyl acetate (3 x 2 mL). The combined organic layer was washed with water (2 x 6 mL), followed by brine (2 x 6 mL). the organic phase was dried over anhydrous Na2SO4 and concentrated in vacuo to afford the crude product, which was purified by FC (n-hexane:ethyl acetate, 7:3), affording the desired product 6 (115 mg, 96%) as a white solid; mp 91-92 oC; [α]20D – 32.4 (c 0.5, CHCl3); 1H NMR (300 MHz, CDCl3) δ 7.51 – 7.24 (m, 11H), 7.21 (d, J = 7.7 Hz, 1H), 7.02 (t, J = 7.4 Hz, 1H), 6.75 (d, J = 7.8 Hz, 1H), 5.32 (d, J = 17.0 Hz, 1H), 5.16 (br s, 1H), 5.00-4.78 (m, 3H), 3.91 – 3.73 (m, 1H), 3.57 – 3.42 (m, 1H), 2.40 (s, 3H), 0.52 (t, J = 7.1 Hz, 3H); 13C NMR (75 MHz, CDCl3) δ 175.9, 165.1, 152.2, 150.7, 142.9, 137.7, 135.6, 131.7, 129.8, 128.9 (2C), 128.7 (2C), 127.8 (3C), 127.1, 126.0 (2C), 123.9, 123.2, 109.1, 101.9, 62.3, 60.0, 46.0, 44.2, 16.8, 13.2; HRMS (ESI) calcd for C29H27N3NaO4+ [MNa]+ 504.1894, found 504.1898. Procedure for the synthesis of (S)-1-benzyl-6'-methyl-2,2'-dioxo-2',3'-dihydro-1'H-spiro[indoline3,4'-pyrimidine]-5'-carboxylic acid (7) Palladium (10 wt.% on carbon, 0.025 mmol, 0.05 eq) was added to a solution of Biginelli-adduct 5j (0.50 mmol, 1 eq) and Et3N (0.50 mmol, 1 eq) in 7.5 mL of dioxane/methanol (2:1). The reaction mixture was degassed in vacuo, placed under an atmosphere of H2 (g), and stirred in the dark at rt for 3h. The mixture was filtered through a pad of Celite eluting with methanol (10 mL), and the combined organic layers were concentrated in vacuo to give the crude carboxylic acid derivative 7 (173 mg, 95%) as a white solid, sufficiently pure to be directly used in the next step; mp not misured (decomposition); [α]20D – 19.2 (c 0.25, CHCl3); 1H NMR (300 MHz, DMSO-d6) δ 11.97 (br s, 1H), 9.39 (br s, 1H), 7.89 (br s, 1H), 7.47 (d, J = 6.7 Hz, 2H), 7.39 – 7.25 (m, 3H), 7.23 (d, J = 7.2 Hz, 1H), 7.16 (t, J = 7.7 Hz, 1H), 6.98 (t, J = 7.4 Hz, 1H), 6.62 (d, J = 7.7 Hz, 1H), 4.96 (d, J = 16.3 Hz, 1H), 4.70 (d, J = 16.3 Hz, 1H), 2.29 (s, 3H); 13C NMR (75 ACS Paragon Plus Environment

The Journal of Organic Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 25

MHz, DMSO-d6) δ 176.1, 166.4, 150.7, 149.4, 142.7, 136.3, 133.9, 128.8, 128.3 (2C), 127.1 (2C), 127.0, 123.0, 122.3, 108.9, 97.8, 63.0, 43.4, 18.5; HRMS (ESI) calcd for C20H17N3NaO4+ [MNa]+ 386.1111, found 386.1121. Procedure for the synthesis of (R)-1-benzyl-6'-methyl-1'H-spiro[indoline-3,4'-pyrimidine]-2,2'(3'H)dione (8) To a solution of the carboxylic acid derivative 7 (0.1 mmol, 1 eq) in 1 mL of dioxane/methanol (1:1), hydrochloric acid in dioxane (4 M, 0.4 mmol, 4 eq) was added, and the reaction was stirred at 90 °C for 0.5h. The solvent was removed under reduced pressure to afford compound 8 (31 mg, 98%) in high purity as a white solid, with no need for further purifications; mp 95-96 oC; [α]20D – 25.6 (c 0.5, CHCl3); 1H NMR (400 MHz, CDCl3) δ 7.82 (br s, 1H), 7.40 (d, J = 7.3 Hz, 1H), 7.37 – 7.23 (m, 5H), 7.18 (t, J = 7.7 Hz, 1H), 7.06 (t, J = 7.5 Hz, 1H), 6.71 (d, J = 7.8 Hz, 1H), 5.72 (br s, 1H), 4.93 (d, J = 15.6 Hz, 1H), 4.79 (d, J = 15.6 Hz, 1H), 4.24 (s, 1H), 1.86 (s, 3H). 13C NMR (100 MHz, CDCl3) δ 177.2, 154.4, 142.0, 136.4, 136.1, 132.6, 130.5, 129.5 (2C), 128.4, 128.0 (2C), 125.7, 124.2, 110.2, 95.4, 64.3, 44.7, 19.4; HRMS (ESI) calcd for C19H17N3NaO2+ [MNa]+ 342.1213, found 342.1206. Procedure

for

the

synthesis

of

diastereoisomers

(S)-1-benzyl-6'-methyl-2,2'-dioxo-N-((S)-1-

phenylethyl)-2',3'-dihydro-1'H-spiro[indoline-3,4'-pyrimidine]-5'-carboxamide

(9a)

and

(R)-1-

benzyl-6'-methyl-2,2'-dioxo-N-((S)-1-phenylethyl)-2',3'-dihydro-1'H-spiro[indoline-3,4'-pyrimidine]5'-carboxamide (9b) To a solution of carboxylic acid derivative 8 (0.9 mmol, 1 eq) and DIPEA (1.8 mmol, 2 eq) in 9.4 mL of anhydrous dimethylformamide, HATU (1.4 mmol, 1.5 eq) was added. After 5 min, (S)-(−)-αmethylbenzylamine (0.9 mmol, 1 eq) and DIPEA (1.8 mmol, 2 eq) were added, and the reaction was stirred at room temperature for 24 hours. The resulting mixture was partitioned between ethyl acetate (20 mL) and water (20 mL). The organic phase was washed with brine (6 x 10 mL), dried over Na2SO4 and concentrated in vacuo to afford the crude diastereoisomeric mixture 9, which was purified by flash chromatography ACS Paragon Plus Environment

Page 21 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

(ethyl acetate:n-hexane, 95:5), obtaining the two isolated stereoisomers 9a (358 mg, 86%) and 9b (54 mg, 12%). 9a: white solid; mp 149-150 oC; [α]20D +16.5 (c 0.9, CHCl3); 1H NMR (400 MHz, DMSO-d6) δ 8.79 (br s, 1H), 8.18 (d, J = 8.4 Hz, 1H), 7.62 (br s, 1H), 7.46 (d, J = 7.3 Hz, 2H), 7.33 (d, J = 7.4 Hz, 1H), 7.32 – 7.22 (m, 7H), 7.18 (q, J = 8.4, 7.8 Hz, 2H), 7.00 (t, J = 7.5 Hz, 1H), 6.63 (d, J = 7.8 Hz, 1H), 4.78 (s, 2H), 4.71 – 4.59 (m, 1H), 1.91 (s, 3H), 1.04 (d, J = 7.0 Hz, 3H). 13C NMR (100 MHz, DMSO-d6) δ 177.4, 165.6, 153.2, 145.3, 144.4, 138.0, 137.3, 132.4, 130.2, 129.4 (2C), 129.3 (2C), 128.3 (2C), 128.1, 127.6, 127.2 (2C), 125.3, 123.0, 109.9, 105.1, 64.3, 48.6, 44.2, 23.0, 18.3. HRMS (ESI) calcd for C28H26N4NaO3+ [MNa]+ 489.1897, found 489.1905. 9b: white solid; mp 138-139 oC; [α]20D – 89.5 (c 1, CHCl3); 1H NMR (400 MHz, DMSO-d6) δ 8.81 (br s, 1H), 8.13 (d, J = 8.3 Hz, 1H), 7.62 (br s, 1H), 7.42 (d, J = 6.4 Hz, 2H), 7.31 (d, J = 7.2 Hz, 1H), 7.29 – 7.19 (m, 3H), 7.17 (t, J = 7.6 Hz, 1H), 7.11 (m, 3H), 7.00 (t, J = 7.4 Hz, 1H), 6.83 (d, J = 7.4 Hz, 2H), 6.57 (d, J = 7.8 Hz, 1H), 4.88 – 4.67 (m, 3H), 2.01 (s, 3H), 1.32 (d, J = 7.0 Hz, 3H). 13C NMR (100 MHz, DMSO-d6) δ 177.4, 165.6, 153.0, 144.9, 144.32, 138.3, 137.3, 132.7, 130.1, 129.4 (2C), 128.9 (2C), 128.2 (2C), 128.07, 127.1, 126.8 (2C), 125.3, 123.2, 110.0, 105.12, 64.4, 48.1, 44.2, 22.5, 18.4. HRMS (ESI) calcd for C28H26N4NaO3+ [MNa]+ 489.1897, found 489.1909.

ASSOCIATED CONTENT Supporting information The Supporting Information is available free of charge on the ACS Publications website at DOI: xxx. 1

H and 13C NMR spectra of all novel compounds, HPLC chromatograms (compounds 5a-j), experimental

for X-ray analysis, computational data.

AUTHOR INFORMATION ACS Paragon Plus Environment

The Journal of Organic Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 25

Corresponding Author * E-mail: [email protected]. Notes The authors declare no competing financial interest.

REFERENCES (1) Singh, G. S.; Desta, Z. Y. Chem. Rev. 2012, 112, 6104-6155. (2) (a) Yang, C.; Li, J.; Zhou, R.; Chen, X.; Gao, Y.; He, Z. Org. Biomol. Chem. 2015, 13,4869–4878. (b) Tian, Y; Nam, S.; Liu, L.; Yakushijin, F.; Yakushijin, K.; Buettner, R.; Liang, W.; Yang, F.; Ma, Y.; Horne, D.; Jove, R. PLoS ONE 2012, 7, e49306. doi:10.1371/journal.pone.0049306. (3) (a) Ruijter, E.; Scheffelaar, R.; Orru, R. V. A. Angew. Chem. Int. Ed. 2011, 50, 6234-6246. (b) van der Heijden, G.; Ruijter, E.; Orru, R. V. A. Synlett 2013, 24, 666–685. (4) Yu, J.; Shi, F.; Gong, L.-Z. Acc. Chem. Res. 2011, 44, 1156-1171. (5) Yu, B.; Yu, D.-Q.; Liu, H.-M. Eur. J. Med. Chem. 2015, 97, 673-698. (6) Liu, L.; Wu, D.; Li, X.; Wang, S.; Li, H.; Li, J.; Wang, W. Chem. Comm. 2012, 48, 1692–1694. (7) Macaev, F.; Sucman, N.; Shepeli, F.; Zveaghintseva, M.; Pogrebnoi, V. Symmetry 2011, 3, 165-170. (8) Dai, W.; Lu, H.; Li, X.; Shi, F.; Tu, S.-J. Chem. Eur. J. 2014, 20, 11382–1389. (9) Shi, F.; Tao, Z.-L.; Luo, S.-W.; Tu, S.-J.; Gong, L.-Z. Chem. Eur. J. 2012, 18, 6885–6894. (10) Tian, L.; Hu, X.-Q.; Li, Y.-H.; Xu, P.-F. Chem. Comm. 2013, 49, 7213-7215. (11) Xie, X.; Peng, C.; He, G.; Leng, H.-J.; Wang, B.; Huang, W.; Han, B. Chem. Comm. 2012, 48, 10487– 10489. (12) (a) Lesma, G.; Meneghetti, F.; Sacchetti, A.; Stucchi, M.; Silvani, A. Belstein J. Org. Chem. 2014, 10, 1383-1389. (b) Sacchetti, A.; Silvani, A.; Gatti, F. G.; Lesma, G.; Pilati, T.; Trucchi, B. Org. Biomol. Chem. 2011, 9, 5515-5522. (c) Lesma, G.; Landoni, N.; Sacchetti, A.; Silvani, A. Tetrahedron 2010, 66, ACS Paragon Plus Environment

Page 23 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

4474-4478. (d) Lesma, G.; Landoni, N.; Pilati, T.; Sacchetti, A.; Silvani, A. J. Org. Chem. 2009, 74, 45374541. (13) Ganem, B. Acc. Chem. Res. 2009, 42, 463-472. (14) Goss, J. M.; Schaus, S. E. J. Org. Chem. 2008, 73, 7651–7656. (15) Heravi, M. M.; Asadi, S.; Lashkariani, B. M. Mol Divers 2013, 17, 389–407. (16) Li, N.; Chen, X.-H.; Song, J.; Luo, S.-W.; Fan, W.; Gong, L.-Z. J. Am. Chem. Soc. 2009, 131, 15301– 15310. (17) Gong, L.-Z.; Chen, X.-H.; Xu, X.-Y. Chem. Eur. J. 2007, 13, 8920–8926. (18) Chen, X.-H.; Xu, X.-Y.; Liu, H.; Cun, L.-F.; Gong, L.-Z. J. Am. Chem. Soc. 2006, 128, 14802-14803. (19) Sohn, J.-H.; Choi, H.-M.; Lee, S.; Joung, S.; Lee, H.-Y. Eur. J. Org. Chem. 2009, 2009, 3858–3862. (20) Xin, J.; Chang, L.; Hou, Z.; Shang, D.; Liu, X.; Feng, X. Chem. Eur. J. 2008, 14, 3177–3181. (21) Wang, Y.; Yang, H.; Yu, J.; Miao, Z.; Chena, R. Adv. Synth. Catal. 2009, 351, 3057–3062. (22) Lahsasni S.; Mohamed M. A. A.; El-Saghier A. M. M. Res. J. Chem. Environ. 2014, 18, 38–42. (23) Kefayati, H.; Rad-Moghadam, K.; Zamani, M.; Hosseyni, S. Lett. Org. Chem. 2010, 7, 277–282. (24) (a) Wallach, O. Liebigs Ann. Chem. 1895, 279, 119−143. (b) Brock, C. P.; Schweizer, W. B.; Dunitz, J. D. J. Am. Chem. Soc. 1991, 113, 9811−9820. (c) Dunitz, J. D.; Gavezzotti, A. J. Phys. Chem. B 2012, 116, 6740−6750. (d) Ernst, K.-H. Phys. Status Solidi B 2012, 249, 2057−2088. (e) Seibel, J.; Parschau, M.; Ernst, K.-H. J. Am. Chem. Soc. 2015, 137, 7970−7973. (f) Mughal, R. K.; Davey, R. J.; Black, S. N. Cryst Growth Des 2007, 7, 225-228. (25) Farrugia, L. J. J. Appl. Cryst. 1999, 32, 837-838. (26) (a) Barone, G.; Gomez-Paloma, L.; Duca, D.; Silvestri, A.; Riccio, R.; Bifulco, G. Chem. Eur. J. 2002, 8, 3233-3239. b) Barone, G.; Gomez-Paloma, L.; Duca, D.; Silvestri, A.; Riccio, R.; Bifulco, G. Chem. Eur. J. 2002, 8, 3240-3245. (27) Pierens, G. K. J. Comp. Chem. 2014, 35, 1388-1394. ACS Paragon Plus Environment

The Journal of Organic Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 25

(28) Smith, S. G.; Goodman, J. M. J. Org. Chem. 2009, 74, 4597-4607. (29) (a) Ma, J. G.; Zhang, J. M.; Jiang, H. H.; Ma, W. Y.; Zhou, J. H. Chin. Chem. Lett. 2008, 19, 375–378. (b) De Souza, R. O. M. A.; da Penha, E. T.; Milagre, H. M. S.; Garden, S. J.; Esteves, P. M.; Eberlin, M. N.; Antunes, O. A. C. Chem. Eur. J. 2009, 15, 9799–9804. (30) Lu, N.; Chen, D.; Zhang, G.; Liu, Q. Int. J. Quantum Chem. 2011, 111, 2031–2038. (31) Folkers, K.; Johnson, T. B. J. Am. Chem. Soc. 1933, 55, 3784–3791. (32) Simón, L.; Goodman, J. M. J. Org. Chem. 2011, 76, 1775–1788. (33) Spartan’08, Wavefunction Inc, Irvine, C.A.; Shao, Y.; Molnar, L.F.; Jung, Y.; Kussmann, J.; Ochsenfeld, C.; Brown, S.T.; Gilbert, A.T.B.; Slipchenko, L.V.; Levchenko, S.V.; O’Neill, D.P.; DiStasio, R.A. Jr; Lochan, R.C.; Wang, T.; Beran, G.J.O.; Besley, N.A.; Herbert, J.M.; Lin, C.Y.; Van Voorhis, T.; Chien, S.H.; Sodt, A.; Steele, R.P.; Rassolov, V.A.; Maslen, P.E.; Korambath, P.P.; Adamson, R.D.; Austin, B.; Baker, J.; Byrd, E.F.C.; Dachsel, H.; Doerksen, R.J.; Dreuw, A.; Dunietz, B.D.; Dutoi, A.D.; Furlani, T.R.; Gwaltney, S.R.; Heyden, A.; Hirata, S.; Hsu, C.-P.; Kedziora, G.; Khalliulin, R.Z.; Klunzinger, P.; Lee, A.M.; Lee, M.S.; Liang, W.Z.; Lotan, I.; Nair, N.; Peters, B.; Proynov, E.I.; Pieniazek, P.A.; Rhee, Y.M.; Ritchie, J.; Rosta, E.; Sherrill, C.D.; Simmonett, A.C.; Subotnik, J.E.; Woodcock, H.L. III; Zhang, W.; Bell, A.T.; Chakraborty, A.K.; Chipman, D.M.; Keil, F.J.; Warshel, A.; Hehre, W.J.; Schaefer, H.F.; Kong, J.; Krylov, A.I.; Gill, P.M.W.; Head-Gordon. M. Phys. Chem. Chem. Phys., 2006, 8, 3172–3191. (34) Lu, S.; BeiPoh, S.; Siau, W.-Y.; Zhao, Y. Angew. Chem. Int. Ed., 2013, 52, 1731-1734. (35) Simonsen, K. B.; Gothelf, K. V.; Jørgensen, K. A. J. Org. Chem., 1998, 63, 7536–7538.

ACS Paragon Plus Environment

Page 25 of 25

The Journal of Organic Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment